You are on page 1of 7

OXIDATIVE ADDITION

Introduction:
Oxidative addition is typically an organometallic chemistry reaction wherein a
molecule adds across a square-planar transition metal complex to form
an octahedral complex. It tends to be paired with reductive elimination, the reverse reaction
if and only if the added ligands were cis. For transition metals, oxidative reaction results in
the decrease in the dn to a configuration with fewer electrons, often 2e fewer. Oxidative
addition is favored for metals that are:
 Basic
 Easily oxidized
Metals with a relatively low oxidation state often satisfy one of these requirements, but
even high oxidation state metals undergo oxidative addition, as illustrated by the oxidation
of Pt(II) with chlorine:
[PtCl4]2− + Cl2 → [PtCl6]2−
The two main observation of oxidative addition reactions are:
 Co-ordination number of central metal atom is increased by +2.
 Oxidation state of metal atom increases by +2.
Oxidative addition requires that the metal complex have a vacant coordination site.
For this reason, oxidative additions are common for four- and five-coordinate complexes.
Explanation:
Oxidative addition is a general method for simultaneously introducing pairs of anionic
ligands, A and B, by the oxidative addition of an A−B molecule such as H2 or CH3‐I.

In the oxidative addition direction, we break the A−B bond and form an M−A and an M−B
bond. The oxidation state (OS), electron count (EC), and coordination number (CN) all increase
by two units during the reaction. It is the change in formal oxidation state (OS) that gives rise
to the oxidative and reductive part of the reaction names.
Oxidative additions proceed by a great variety of mechanisms, however, a vacant 2e
site is always required on the metal. We can either start with a 16e complex or a 2e site must
be opened up in an 18e complex by the loss of a ligand producing a 16e intermediate species.
The change in oxidation state means that the starting metal complex of a given oxidation state
must also have a stable oxidation state two units higher to undergo oxidative addition (and
vice versa for reductive elimination).
The reverse reaction, reductive elimination, leads to the extrusion of A−B from an
M(A)(B) complex and is often the product‐forming step in a catalytic reaction.
Binuclear Oxidative Addition:
Each of two metals change their oxidation states, electron count, and coordination
number by one unit each.
This typically occurs in the case of a 17e complex or a binuclear 18e complex with an
M−M
Bond where the metal has a stable oxidation state more positive by one unit. Whatever the
mechanism, there is a net transfer of two electrons into the σ􀗛 orbital of the A−B bond, and
the two A−B σ electrons are divided between both metals. This cleaves the A−B bond and
makes and M−A and M−B bond.
As we have seen, oxidative addition is the inverse of reductive elimination and vice
versa. In principle, these reactions are reversible, but in practice they tend to go in the
oxidative or reductive direction only. The position of equilibrium in any particular case is
governed by the overall thermodynamics and relative stabilities of the two oxidation states.
Alkyl hydride complexes commonly undergo reductive elimination of an alkane, but
rarely does oxidative addition of alkanes occur. Conversely, alkyl halides commonly undergo
oxidative addition, but the adducts rarely reductively eliminate the alkyl halide. Rare
examples of equilibrium do exist, but are thermodynamically controlled.
Pathways for Oxidative Addition:
Oxidative additions are very diverse mechanistically, and we therefore consider each
type separately.
1. Concerted, or three‐center, oxidative addition mechanism
2. SN2 mechanism
3. Radical mechanisms
4. Ionic Mechanisms
Concerted Mechanism
Concerted, or three‐center, oxidative addition is really an associative reaction in which
the incoming ligand first binds as a σ complex and then undergoes bond breaking as a result
of
strong back donation g from the metal into the σ* orbital. Non‐polar reagents, such as H2, or
compounds containing C−H and Si−H bonds all tend to react via a σ complex transition state
(or even an intermediate) of this type. The associative step a involves formation of a σ
complex; sometimes this is stable and the reaction stops here. Step b is the oxidative part of
the reaction in which metal electrons are formally transferred to the σ* orbital of A−B.
There are many examples, however, one of the most‐studied cases is the addition of
H2 to the 16e square planar d8 species IrCl(CO)(PPh3)2 [aka Vaska’s complex ] to give the 18e
d6
3 octahedral dihydride IrCl(H2)(CO)(PPh3)2 .

Normally two ligands that are trans in the Ir(I) complex fold back to give the cis
dihydride isomer, but subsequent rearrangement can occur. Conversely, in a reductive
elimination such as the loss of H2 from the dihydride, the two ligands to be eliminated
normally have to be cis to one another.
Oxidative addition of oxygen also takes place through concerted mechanism:

SN2 Mechanism:
In all oxidative additions, a pair of electrons from the metal is used to break the A−B
bond in the reagent. In the SN2 pathway, adopted for polarized A‐B substrates such as alkyl
halides, the metal electron pair of LnM directly attacks the A–B σ􀗛 orbital by an in‐line attack
at the least electronegative atom (where σ􀗛 is largest) formally to give LnM2+ , A−, and B−
fragments (ionic model). The SN2 mechanism is often found in the addition of methyl, allyl,
acyl, and benzyl halides. Like the concerted type, they are second‐order reactions, but they
are acceleration polar solvents and show negative entropies of activation. This is consistent
with an ordered, polar transition state, as in organic SN2 reactions.

R and X may end up cis or trans to one another in the final product, as expected for
the recombination of the ion pair formed in the first step.
Of the two steps,
 the first involves oxidation by two units but no change in the electron count (Me+ is a
0e reagent)

 The second step involves an increase by 2e in the electron count (I− is a 2e reagent)
but no change in the oxidation state.

Only the two steps together constitute the full oxidative addition. When an 18e complex is
involved, the first step can therefore proceed without the necessity of losing a ligand first.
Only the second step requires a vacant 2e site.
 The more nucleophilic the metal, the greater its reactivity in SN2 additions, as
illustrated by the reactivity order for some Ni(0) complexes:
Ni(PR3)4 > Ni(PAr3)4 > Ni(PR3)2(alkene) > Ni(PAr3)2(alkene) > Ni(cod)2 (R = alkyl)
 Steric hindrance at carbon slows the reaction, so we find the reactivity order:
Me‐I > Et‐I > iPr‐I
 A better leaving group accelerates the reaction, which gives rise to the reactivity
order:
R‐OSO2(C6H4Me) > R‐I > R‐Br > R‐Cl

Radical mechanism
Radical mechanisms in oxidative additions were recognized later than the SN2 and the
concerted processes. They can also be photo initiated. A troublesome feature of these
reactions is that minor changes in the structure of the substrate, the complex, or in impurities
present in the reagents of solvents can sometimes be enough to change the rate, and even
the predominant mechanism of a given reaction. Sharp disagreements have turned on
questions of repeatability and on what types of experimental evidence should be considered
as valid mechanistic criteria. For example, the use of radical traps, such as RNO•, has been
criticized on the grounds that these may initiate a radical pathway for a reaction that
otherwise would have followed a non‐radical mechanism in the absence of trap.
Two subtypes of radical process are now distinguished:
 Non‐chain
 Chain
The non‐chain variant is believed to operate in the additions of certain alkyl halides,
R‐X, to Pt(PPh3)3 (R = Me, Et; X = I; R = PhCH2; X = Br).
A chain process occurs if the radicals formed escape from the solvent cage without
recombination. Otherwise, a radical initiator, Q•, (e.g., a trace of air) may be required to set
the process going. This can lead to an induction period (a period of dead time before the
reaction starts).

Ionic mechanism
Hydrogen halides are often largely dissociated in solution, and the anion and proton
tend to add to metal complexes in separate steps. Two variants have been recognized.
In the more common one, the complex is basic enough to protonate, after which the
anion binds to give the final product.
 protonation – anionation
Rarer is the opposite case in which the halide ion attacks first, followed by protonation of
the intermediate.
 Anionation ‐ protonation
In the more common mechanism of protonation followed by anionation, the
complex is basic enough to protonate, after which the anion binds to give the final product.
This route is favored by
 Polar solvents
 Basic ligands
 A low‐oxidation‐state metal.
 Synthetic Applications of oxidative Addition of O2
Addition of Oxygen to Alkenes

Addition of oxygen to alkenes follow through ozonolysis process which results in


hydroformylation of alkene and forms aldehydes and ketones depending upon the
constituent alkene. This reaction is catalyzed by transition metal i.e. Zn. The importance and
relevance of this reaction to modern synthetic chemistry is great. With the CHO group a new
functional group is created and immensely important class of organic compounds is
accessible. Similarly formyl groups can be easily converted to amines, imines, hemiacetal,
acetals, and aminals and so on.

1. Ziegler-Natta Catalysis
A Ziegler–Natta catalyst, named after Karl Ziegler and Giulio Natta, is a catalyst used in the
synthesis of polymers of 1-alkenes (alpha-olefins). These catalysts are used to polymerize
terminal alkenes (ethylene and alkenes with the vinyl double bond):
n CH2=CHR → −[CH2−CHR]n−;
References:
alpha.chem.umb.edu/.../Lec16OxidativeAddition_000.pdf
web.iitd.ac.in/~sdeep/Elias_Inorg_lec_7.pdf
en.wikipedia.org/wiki/Oxidative_addition

You might also like